首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 203 毫秒
1.
A method of electron paramagnetic resonance was applied to the study of a possibility of paramagnetic centres formation in the nitrogen oxide reaction with saturated and unsaturated fatty acids and aromatic amino acids. Several paramagnetic centres are formed in interaction of NO2 with oleic acid at 20 degrees C. Saturated fatty acids formed no paramagnetic centres in the reaction with NO2. NO formed no paramagnetic centres not only with butyric, palmitic, and stearic acids, but also with oleic acid. The capacity of NO2 to form paramagentic centres with tyrosine in its saturated solution was revealed. In case of interaction between NO and NO2 with saturated phenylalanine and tryptophane solutions no formation of paramagnetic centres was observed. It is suggested that unsaturated fatty acids and tyrosine remnants of the membrane lipoprotein complexes could serve as peculiar targets with which NO2 interacted, and that formation of paramagnetic centres in the NO2 reaction with the mentioned substances was the primary act of the membrane damage.  相似文献   

2.
Acholeplasma laidlawii, a mycoplasma, is unable to synthesize unsaturated fatty acids but it will incorporate them into its plasma membrane if they are supplied exogeneously. Thus the fatty acid composition of the cell membrane can be defined by growing the organism in media containing specific fatty acids. We obtained cells with predominantly one type of unsaturated fatty acid (either oleic, linoleic or linolenic acid) or cells with only saturated fatty acid in the cell membrane. The cells were irradiated with 7 MeV electrons and the effect of membrane fatty acid composition on cell survival was examined. At 200 Gy/min and 0.5 degrees C (melting ice) there was little difference in the radiation sensitivities of the cells grown in unsaturated fatty acids either in aerated or anoxic radiation conditions. However, the cells containing saturated fatty acids irradiated in anoxic conditions were markedly more sensitive than the cells containing unsaturated fatty acids. At 200 Gy/min and 37 degrees C the two types of cells were of similar sensitivity both in aerated and anoxic radiation conditions. At 5 Gy/min at 0.5 degrees C the cells containing linolenic acid (18:3) were less sensitive than those containing solely saturated fatty acids. However, at 5 Gy/min at 37 degrees C there was no difference in sensitivity between these two types of cell. Our results strongly argue against the involvement of lipid peroxidation as a molecular change leading to cell death.  相似文献   

3.
Non-esterified long-chain fatty acids reduce the extent of hypotonic hemolysis at a certain low concentration range but cause hemolysis at higher concentrations. This biphasic behavior was investigated at different temperatures (0-37 degrees C) for lauric (12:0), myristic (14:0), palmitoleic (16:1), oleic (cis-18:1) and elaidic (trans-18:1) acids. The results are summarized as follows: (A) the fatty acids examined exhibit a high degree of specificity in their thermotropic behavior; (B) oleic acid protects against hypotonic hemolysis even at the highest concentrations, up to 15 degrees C, when it becomes hemolytic, but only in a limited concentration range; (C) elaidic acid does not affect the osmotic stability of erythrocytes up to 20 degrees C, when it starts protecting: above 30 degrees C, it becomes hemolytic at the highest concentrations; (D) palmitoleic acid is an excellent protecting agent at all temperatures in a certain concentration range, becoming hemolytic at higher concentrations; (E) lauric acid protects up to 30 degrees C and becomes hemolytic only above this temperature; (F) myristic acid exhibits an extremely unusual behavior at 30 and 37 degrees C by having alternating concentration ranges of protecting and hemolytic effects; (G) there is a common critical temperature for hemolysis at 30 degrees C for saturated and trans-unsaturated fatty acids; (H) the initial slope of Arrhenius plots of percent hemolysis at the concentration of maximum protection is negative for cis-unsaturated fatty acids and positive for saturated and trans-unsaturated fatty acids.  相似文献   

4.
When Streptococcus salivarius was grown in batch culture in the presence of various Tween detergents, the fatty acid moiety of the detergent was incorporated into the lipids of its membrane. Tween 80 (containing primarily oleic acid) markedly stimulated the production of extracellular glucosyltransferase and also increased the degree of unsaturation of the membrane lipid fatty acids. The possibility that an increase in membrane unsaturated fatty acids promoted extracellular glucosyltransferase production was examined by growing cells at different temperatures in the presence or absence of Tween 80. The membrane lipids of cells grown at 30 degrees C, 37 degrees C and 40 degrees C without Tween 80 exhibited unsaturated/saturated fatty acid ratios of 2.06, 1.01 and 0.87 respectively. A significant increase in the production of extracellular glucosyltransferase was observed at 30 degrees C compared to cells grown at 40 degrees C. However, cells produced much more exoenzyme at all temperatures when grown with Tween 80. The results indicated that an increase in the unsaturated fatty acid content of the membrane lipids was not by itself sufficient to account for the stimulation of extracellular glucosyltransferase production by Tween 80, but that the surfactant also had to be present.  相似文献   

5.
Mycobacterium smegmatis ATCC 607 was grown at 27 and 37°C, with and without exogenous unsaturated fatty acids, viz. elaidic, oleic and palmitoleic acids, added to the growth medium. The total lipid content of M. smegmatis ATCC 607 was lower at 27°C, and with added oleic acid, when compared with the controls, but higher in presence of palmitoleic acid. At 37°C no significant differences were noted in the total lipid content. In general, the total lipid content was lower with all of the fatty acid supplementations at both 27 and 37°C. The phosphatidylethanolamine content was slightly higher at 27°C in the presence of elaidic or palmitoleic acid, but was markedly lower with oleic acid supplementation at 37°C. The cardiolipin content was lower in the presence of any of the fatty acids at 27°C, and higher in the medium supplemented with elaidic or oleic acid at 37°C. The unsaturated to saturated fatty acids ratio was higher with palmitoleic acid supplementation at 27°C, but remained unchanged in cells grown at 37°C. The modifications in mycobacterial lipids are a reflection of the organism's ability to adapt to changing growth conditions.  相似文献   

6.
The effects of various non-esterified fatty acids on the CETP-mediated particle size redistribution of HDL were studied by incubating HDL3 and CETP for 24 h at 37 degrees C in the absence or in the presence of either saturated, monounsaturated or polyunsaturated non-esterified fatty acids. In the absence of non-esterified fatty acids, CETP induced a redistribution of the initial population of HDL3 (Stokes' radius 4.3 nm) by promoting the appearance of one larger (Stokes' radius 4.8 nm) and two smaller (Stokes' radii 3.9 and 3.7 nm) HDL subpopulations. Whereas the non-esterified fatty acids alone did not modify the HDL3 distribution profile, they were able to alter markedly the capacity of CETP to induce the particle size redistribution of HDL. All the saturated fatty acids with at least 10 carbons were able to increase the formation of the very small sized particles (Stokes' radius 3.7 nm) in a concentration dependent manner, the medium chain fatty acids (12 and 14 carbons) being the best activators. The potential effect of non-esterified fatty acids was also influenced by the presence of double bonds in their monomeric carbon chain. While at low concentrations of non-esterified fatty acids (0.1 mmol/l) the enhancement of the formation of very small HDL particles appeared to be greater with oleic and linoleic acids than with stearic acid, at higher concentrations (0.4 mmol/l), oleic, linoleic and arachidonic acids decreased the formation of the 3.7 nm radius particles. The inhibition of the process at high concentrations of unsaturated fatty acids was linked to the degree of unsaturation of their carbon chain, arachidonic acid being the strongest inhibitor. The present study has demonstrated that non-esterified fatty acids can modulate the particle size redistribution of HDL3 mediated by the cholesteryl ester transfer protein even in the absence of any other lipoprotein classes. The effect of non-esterified fatty acid is dependent on both the length and the degree of unsaturation of their monomeric carbon chain.  相似文献   

7.
Fatty acids newly synthesized by Brevibacterium ammoniagenes grown at different temperatures were analyzed. The assay temperature, not the growth temperature, was found to be the major factor affecting the unsaturated/saturated ratio of newly synthesized fatty acids in logarithmic-phase cells. However, in the stationary-phase cells the growth temperature also affected the product profile significantly; cells grown at 7 degrees C produced relatively more oleate and stearate and less palmitate and hexadecenoate when shifted up to 37 degrees C than did cells grown and assayed at 37 degrees C. The unsaturated/saturated ratio as well as average chain length of fatty acids also varied along with the progress of isothermal growth phase. These changes in fatty acid product profiles observed in vivo could be mimicked in vitro assays of the fatty acid synthetase by changing malonyl-CoA concentrations. Our results suggest that the malonyl-CoA concentration is a factor which, in addition to temperature, determines growth-phase-dependent and growth-temperature-dependent changes in the unsaturated/saturated ratios of fatty acids.  相似文献   

8.
A conditional (temperature sensitive) fatty acid biosynthetic mutant (fabB2) of Salmonella typhimurium does not support the development of the virulent bacteriophage 9NA even at permissive temperature (30 degrees C). A limited amount of phage DNA synthesis takes place at this temperature. When the fatty acid composition of the host membrane is altered by growing the cells at 37 degrees C in the presence of exogenous unsaturated fatty acid, differential expression of phage genes was observed. Phage specific lysozyme is induced when the cultures are supplemented with elaidic, palmitelaidic, linoleic and linolelaidic acids but not with oleic and plamitoleic acids. However, in no case were infective particles produced. Under conditions where no lysozyme is synthesized the infected cells increase in length and become filamentous.  相似文献   

9.
The effect of long-chain fatty acids on the binding of thyroxine to highly purified human thyroxine-binding globulin has been studied by equilibrium dialysis performed at pH 7.4 and 37 degrees C. At a fixed molar ratio of 2000:1 of fatty acid to thyroxine-binding globulin, the degree of binding inhibition based on the percent change in nK value relative to the control as determined from Scatchard plots was: palmitic, 0%; stearic, 0%; oleic, 76%; linoleic, 69%; and linolenic, 61%. At a 500:1 molar ratio of oleic acid to thyroxine-binding globulin, equivalent to 0.125 mM free fatty acid in serum, thyroxine binding was inhibited by 18%, increasing to 93% at a 4500:1 molar ratio. At molar ratios of oleic acid to thyroxine-binding globulin of 1000:1, 2000:1 and 4000:1, the degree of inhibition of triiodothyronine binding was 24%, 41% and 76%, respectively. The results indicate that the unsaturated long-chain fatty acids are potent inhibitors of thyroxine binding to thyroxine-binding globulin, whereas the saturated fatty acids have little or no effect on thyroxine binding.  相似文献   

10.
Three of the major phospholipids of the cell membrane of Clostridium butyricum are phosphatidylethanolamine (PE), plasmenylethanolamine (PlaE), and the glycerol acetal of plasmenylethanolamine. When cultured in the absence of biotin in media supplemented with a cis-unsaturated fatty acid, the cellular lipids become highly enriched with the fed fatty acid. Under these conditions, the ratio of the glycerol acetal of PlaE to the sum of PE plus PlaE increases markedly over that seen in cells containing mixtures of saturated and unsaturated fatty acids [Johnston, N.C., & Goldfine, H. (1985) Biochim. Biophys. Acta 813, 10-18]. We have studied the polymorphic phase behavior of the phospholipids from C. butyricum grown on oleic acid using differential scanning calorimetry, 31P nuclear magnetic resonance, and X-ray diffraction. The mixed PE plus PlaE fraction undergoes a transition from the gel to liquid-crystalline state at -1.9 degrees C and a lamellar to reversed hexagonal (L----H) transition at or near 0 degrees C. The glycerol acetal of PlaE melts at 16.1 degrees C, and as predicted from lipid packing theory, the lamellar phase is stabilized, up to 50 degrees C. Addition of the oleate-enriched glycerol acetal of PlaE to dioleoylphosphatidylethanolamine, or the PE plus PlaE fraction from oleate-grown cells, stabilized the lamellar arrangement of the mixtures. A ratio of glycerol acetal of PlaE to total PE (PE plus PlaE) of 0.5, which is close to that found in cells grown on palmitic plus oleic acid, 0.6-0.7, did not produce a lamellar phase at 37 degrees C when the lipids enriched with oleic acid were tested,(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Structural analysis of phosphatidylcholine of plant tissue   总被引:3,自引:0,他引:3  
Pure preparations of phosphatidylcholine were isolated from spinach leaf chloroplasts, spinach leaf microsomes, and cauliflower inflorescence. The isolated phosphatidylcholine was treated with snake venom phospholipase A, and the fatty acid distribution and composition of the fatty acid methyl esters prepared from the lysophosphatidylcholine and the freed fatty acid were determined by gas-liquid chromatography. The results showed that saturated fatty acids were preferentially esterified at position 1 and unsaturated fatty acids at position 2. The phosphatidylcholine from cauliflower was also treated with phospholipase C. The resulting diglycerides were fractionated on AgNO(3)-impregnated thin-layer plates. The diglyceride fractions were transesterified and the fatty acid composition of each was determined by gas-liquid chromatography. The predominant species contained linolenic acid only (22% of the total), linolenic and oleic acids (19%), and linolenic and palmitic acids (37%). These molecular species could not be accounted for by random distribution of the fatty acids.  相似文献   

12.
The addition of 0.4-3 mM of cis-unsaturated fatty acids such as oleic acid (18:1) or linoleic acid (18:2) to intact rat adipocytes stimulated lipogenesis at 37 degrees C. Saturated or trans-unsaturated fatty acids were ineffective. Fluorescence photobleaching recovery studies performed under similar conditions indicated that the cis-unsaturated fatty acids do not alter lateral mobility of either a lipid probe or a general protein marker in the plasma membrane. A high concentration (7 mM) of Ca2+, which by itself has some stimulatory effect on lipogenesis, significantly potentiated the effect of oleic acid on this insulin-like activity. Measurement of 45Ca2+ binding by fat cells has indicated that cis-unsaturated (but not saturated) fatty acids increased 12- to 20-fold the amount of Ca2+ associated with the cells. The dependence of this effect on the fatty acid concentration correlates well with the effect of the fatty acid on the induction of lipogenesis. Our results suggest that cis-unsaturated fatty acids affect membrane organization in a manner which induces a significant increase in membrane associated or intracellular Ca2+. This increase may be responsible for inducing exocytotic-like processes which facilitate translocation of glucose transport activity from storage sites to the plasma membrane and thus produce an insulin-like effect.  相似文献   

13.
Unsaturated fatty acids are important constituents of all cell membranes and are required for normal growth. In the filamentous fungus Aspergillus nidulans, unsaturated fatty acids and their derivatives also influence asexual (conidial) and sexual (ascospore) sporulation processes. To investigate the relationship between fatty acid metabolism and fungal development, we disrupted the A. nidulans sdeA and sdeB genes, both encoding Delta9-stearic acid desaturases responsible for the conversion of palmitic acid (16:0) and stearic acid (18:0) to palmitoleic acid (16:1) and oleic acid (18:1). The effects of sdeA deletion on development were profound, such that growth, conidial and ascospore production were all reduced at 22 and 37 degrees C. Total fatty acid content was increased over 3-fold in the DeltasdeA strain, reflected in up-regulation of the expression of the fasA gene encoding the alpha chain of the fatty acid synthase, compared to wild type. Stearic acid accumulated approximately 3-fold compared to wild type in the DeltasdeA strain, while unsaturated fatty acid production was decreased. In contrast, disruption of sdeB reduced fungal growth and conidiation at 22 degrees C, but did not affect these processes at 37 degrees C compared to wild type. Interestingly, ascospore production was increased at 37 degrees C for DeltasdeB compared to wild type. Total fatty acid content was not increased in this strain, although stearic acid accumulated 2-fold compared to wild type, and unsaturated fatty acid production was decreased. Combining the DeltasdeA and DeltasdeB alleles created a synthetic lethal strain requiring the addition of oleic acid to the medium for a modicum of growth. Taken together, our results suggest a role for sdeA in growth and development at all temperatures, while sdeB is involved in growth and development at lower temperatures.  相似文献   

14.
We measured the influence of saturated and unsaturated free fatty acids on the permeability and partition of ions into 1, 2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) bilayers. The bilayer permeability was measured using the depletion of N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1, 2-dihexadecanoyl-sn-glycero-3-phosphatidylethanolamine (N-NBD-PE) fluorescence as a result of its reduction by dithionite. We observed a distinct increase of dithionite permeability at the main gel-fluid phase transition of DMPC. When vesicles were formed from a mixture of DMPC and oleic acid, the membrane permeability at the phase transition was reduced drastically. Stearic acid and methyl ester of oleic acid have little effect. Similar results in the quenching of pyrene-PC in DMPC vesicles by iodide were obtained. Again, the increase of iodide partition into the lipid phase at the main phase transition of DMPC was abolished by the addition of unsaturated free fatty acids. Free fatty acids, in concentrations up to 5 mol%, do not abolish DMPC phase transition when measured by differential scanning calorimetry. It seems that unsaturated, but not saturated, free fatty acids reduce the lipid bilayer permeability to dithionite and iodide ions at the main phase transition of DMPC, without altering the thermodynamic properties of the bilayer.  相似文献   

15.
Livers of fasted rats were perfused for 70 min at 37 degrees-43 degrees C in the presence or absence of acetate, octanoate or palmitate. Hepatic biosynthetic capacity was assessed by measuring rates of gluconeogenesis, ureogenesis, ketogenesis and O2 consumption. In the presence of each fatty acid, gluconeogenesis, ureogenesis and oxygen consumption were maintained at 37 degrees and 42 degrees C. At 43 degrees, the rate of glucose formation decreased markedly and rates of ureogenesis and oxygen consumption were distinctly lower. As the temperature was increased from 37 degrees to 43 degrees C without fatty acids, i.e. albumin only, there was a progressive decrease in the rate of gluconeogenesis while the ratio of net C3 utilized to glucose formed, increased successively. The values of this ratio in the presence of palmitate or octanoate at 43 degrees were smaller than those for albumin or acetate, but higher than the figure of 2 for complete conversion of C3 units to glucose. Although fatty acid was added in equimolar amounts of C2 units, total ketone formation was influenced significantly by chain length. Hepatic ketogenesis was similar at 37 degrees with albumin, palmitate, or acetate, but was stimulated significantly by octanoate at 37 degrees and 42 degrees C. At 42 degrees, ketone formation increased in the presence of palmitate. At 43 degrees C, ketogenesis with palmitate or octanoate decreased, while that with acetate or albumin was maintained at the same lower rates. The ratio of 3-hydroxybutyrate to acetoacetate in the perfusate was increased with palmitate at the end of perfusion at 37 degrees and 42 degrees C or octanoate at 42 degrees and 43 degrees C. Thus, long (palmitate)- and medium (octanoate)- but not short (acetate)-chain fatty acids enhance not only beta-oxidation, but influence the redox state of hepatic mitochondria with an increase in the state of reduction of the pyridine nucleotides. Such a shift in the redox state would be operable in the perfused liver even at 43 degrees C and may be responsible for improved conversion of lactate to glucose when medium- or long-chain fatty acids are present at hyperthermic temperatures.  相似文献   

16.
In a previous publication (Lagrost, L. and Barter, P.J. (1991) Biochim. Biophys. Acta 1085, 209-216), saturated and cis unsaturated non-esterified fatty acids have been shown to modulate the rate at which cholesteryl esters are transferred from high-density lipoproteins (HDL) to low-density lipoproteins (LDL) in the presence of the human cholesteryl ester transfer protein (CETP). In the present report, the effects of cis (oleic acid) and trans (elaidic acid) monounsaturated isomers on the CETP-mediated transfer of cholesteryl esters between HDL and LDL were compared. Mixtures of human LDL and HDL3, containing or not radiolabelled cholesteryl esters, were incubated at 37 degrees C with CETP in the presence or in the absence of either stearic (18:0), oleic (18:1 cis) or elaidic (18:1 trans) acids. It was observed that oleic acid and elaidic acid had different effects on the CETP-mediated redistribution of radiolabelled cholesteryl esters as well as on the net mass transfer of cholesterol from HDL3 to LDL. In particular, at high non-esterified fatty acid/lipoprotein ratio, the transfer of cholesteryl esters was significantly inhibited by the cis isomer and increased by the trans isomer.  相似文献   

17.
Microbial biohydrogenation of oleic acid to trans isomers in vitro   总被引:5,自引:0,他引:5  
Ruminant products are significant sources of dietary trans fatty acids. Trans fatty acids, including various conjugated linoleic acid isomers, have been shown to act as metabolic modifiers of lipid metabolism. Trans fatty acids originate from biohydrogenation of dietary unsaturated fatty acids by gut microbes; however, the exact synthetic pathways are unclear. It was our goal to examine the biohydrogenation pathway for oleic acid, where oleic acid is hydrogenated directly to stearic acid. Our objective in this study was to trace the time course of appearance of 13C in labeled oleic acid to determine if trans monoenes are formed from the 13C-labeled oleic acid or if the 13C appears only in stearic acid as described in reviews of earlier work. Enrichments were calculated from the mass abundance of 13C in major fatty acid fragments and expressed as a percentage of total carbon isotopomers. Significant 13C enrichment was found in stearic acid, oleic acid, trans-6, trans-7, and in all trans C18:1 in positions 9-16. We concluded that the biohydrogenation of oleic acid by mixed ruminal microbes involves the formation of several positional isomers of trans monoenes rather than only direct biohydrogenation to form stearic acid as previously described.  相似文献   

18.
The effects of unsaturated fatty acids on the activities of peroxisomal enzymes of Tetrahymena pyriformis were investigated. When saturated fatty acids and the corresponding unsaturated fatty acids (C18) were added to the culture medium at 0.05%, the activities of peroxisomal enzymes [fatty acyl-CoA oxidase (FAO), carnitine acetyltransferase (CAT), isocitrate lyase (ICL), and malate synthase (MS)] were significantly increased. The order of effectiveness was linoleic acid greater than oleic acid greater than stearic acid. However, alpha-linolenic acid and gamma-linolenic acid at the same concentration were lethal to the cells. The inhibitory effect on growth disappeared upon addition of an antioxidant, alpha-tocopherol. Lipid peroxides derived from unsaturated fatty acids induced marked cell lysis. In the presence of a low concentration (0.005%) of linolenic acid the production of lipid peroxide was lower and no inhibitory effect on the growth was observed, while the activities of peroxisomal enzymes participating in lipid metabolism and that of catalase were significantly increased. These results indicate that the peroxisomal enzyme systems related to the beta-oxidations of fatty acids and the glyoxylate cycle are regulated by unsaturated long-chain fatty acids, including linolenic acid, at low concentrations, as well as by saturated fatty acid in the medium.  相似文献   

19.
The effect of various fatty acids on lipid peroxidation of liver microsomes induced by different methods in vitro was studied using oxygen uptake and malonaldehyde (MDA) production. It was observed that fatty acids with a single double bond are effective inhibitors of peroxidation. Stereo and positional isomers of oleic acid were equally effective as oleic acid. There was an absolute requirement for a free carboxyl group, since methyl esters of fatty acids and long-chain saturated and unsaturated hydrocarbons could not inhibit peroxidation. Saturated fatty acids with a chain length of 12-16 carbon atoms showed inhibition, whereas more than 18 carbon atoms reduced the inhibitory capacity. Fatty acids of lower chain length such as capric and caprylic acids did not show inhibition. Fatty acid inhibition was partially reversed by increasing the concentration of iron in the system. Peroxidation induced by methods which were independent of iron was not inhibited by fatty acids. It was observed that intestinal microsomes which were resistant to peroxidation due to the presence of nonesterified fatty acids in their membrane lipids were able to peroxidise by methods which do not require iron. These results suggest that certain fatty acids inhibit peroxidation by chelating available free iron. In addition, they may also be involved in competing with the esterified fatty acids in the membrane lipids which are the substrates for peroxidation.  相似文献   

20.
We examined effects on intestinal absorption of cholesterol and triglycerides and intestinal lipoprotein formation by feeding rats diets in which saturated fatty acids (palmitic plus stearic) comprised 78%, 68%, 48%, or 38% of triglyceride fatty acids. Absorption into lymph of radiolabeled cholesterol was proportional to triglyceride absorption. The rates of absorption of these lipids were related inversely to the % saturated fatty acids fed. The distribution of newly absorbed cholesterol and triglyceride into intestinal lipoproteins differed. With increasing cholesterol absorption more was recovered in very low density lipoproteins in contrast to the appearance preferentially in chylomicrons of larger quantities of fatty acid. Lymph lipid content did not reflect a consistent pattern in relation to the experimental diet fed. The fatty acid composition of triglyceride-rich lymph lipoproteins resembled the diet closely. One-quarter of the intestinal lymph particles from rats fed the highly saturated diets was flattened and polygonal as judged by electron microscopy if cooled to room temperature; whereas with the same diets, particles collected and isolated at 37 degrees C were round. Proportions of A-I and C apolipoproteins in triglyceride-rich intestinal particles varied inversely; apoA-I increased as fat/cholesterol absorption was greater. Diet-induced alterations in plasma lipoproteins and increased circulating triglycerides in this study in rats were unrelated to the variations in intestinal absorption or lymph lipoprotein formation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号